arxiv: v1 [cond-mat.stat-mech] 14 Apr 2016

Similar documents
arxiv: v4 [cond-mat.stat-mech] 17 Dec 2015

Fluctuation theorem between non-equilibrium states in an RC circuit

Fluctuation response inequality out of equilibrium

Stochastic thermodynamics

Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences

Contrasting measures of irreversibility in stochastic and deterministic dynamics

Introduction to Fluctuation Theorems

On the Asymptotic Convergence. of the Transient and Steady State Fluctuation Theorems. Gary Ayton and Denis J. Evans. Research School Of Chemistry

Fluctuation theorem in systems in contact with different heath baths: theory and experiments.

LANGEVIN EQUATION AND THERMODYNAMICS

Non-equilibrium phenomena and fluctuation relations

Nonequilibrium thermodynamics at the microscale

Fluctuation theorems. Proseminar in theoretical physics Vincent Beaud ETH Zürich May 11th 2009

J. Stat. Mech. (2011) P07008

Entropy production and time asymmetry in nonequilibrium fluctuations

arxiv: v2 [cond-mat.stat-mech] 16 Mar 2012

CHAPTER V. Brownian motion. V.1 Langevin dynamics

arxiv: v2 [cond-mat.stat-mech] 28 Mar 2013

Mathematical Structures of Statistical Mechanics: from equilibrium to nonequilibrium and beyond Hao Ge

Anomalous Transport and Fluctuation Relations: From Theory to Biology

Fluctuation theorems: where do we go from here?

Thermodynamics for small devices: From fluctuation relations to stochastic efficiencies. Massimiliano Esposito

Beyond the Second Law of Thermodynamics

Information Thermodynamics on Causal Networks

The physics of information: from Maxwell s demon to Landauer. Eric Lutz University of Erlangen-Nürnberg

The Kramers problem and first passage times.

arxiv: v3 [cond-mat.stat-mech] 15 Nov 2017

The dynamics of small particles whose size is roughly 1 µmt or. smaller, in a fluid at room temperature, is extremely erratic, and is

arxiv: v1 [cond-mat.stat-mech] 6 Mar 2008

arxiv: v2 [cond-mat.stat-mech] 3 Mar 2018

Stochastic equations for thermodynamics

arxiv: v1 [cond-mat.stat-mech] 2 Jul 2008

Statistical Mechanics of Active Matter

arxiv: v1 [cond-mat.stat-mech] 15 Sep 2007

The Kawasaki Identity and the Fluctuation Theorem

Anomalous Lévy diffusion: From the flight of an albatross to optical lattices. Eric Lutz Abteilung für Quantenphysik, Universität Ulm

NPTEL

Statistical properties of entropy production derived from fluctuation theorems

Second law, entropy production, and reversibility in thermodynamics of information

Diffusion in the cell

COARSE-GRAINING AND THERMODYNAMICS IN FAR-FROM-EQUILIBRIUM SYSTEMS

This is a Gaussian probability centered around m = 0 (the most probable and mean position is the origin) and the mean square displacement m 2 = n,or

Molecular Machines and Enzymes

Langevin Methods. Burkhard Dünweg Max Planck Institute for Polymer Research Ackermannweg 10 D Mainz Germany

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Statistical Physics I Spring Term 2013 Notes on the Microcanonical Ensemble

Statistical Mechanics and Thermodynamics of Small Systems

Brownian Motion. 1 Definition Brownian Motion Wiener measure... 3

Optimum protocol for fast-switching free-energy calculations

Brownian Motion and Langevin Equations

Entropy Production and Fluctuation Relations in NonMarkovian Systems

Introduction to nonequilibrium physics

MD Thermodynamics. Lecture 12 3/26/18. Harvard SEAS AP 275 Atomistic Modeling of Materials Boris Kozinsky

arxiv: v4 [cond-mat.stat-mech] 3 Mar 2017

Information Theory in Statistical Mechanics: Equilibrium and Beyond... Benjamin Good

Brownian motion and the Central Limit Theorem

Fluctuation Theorems of Work and Entropy in Hamiltonian Systems

Stochastic Processes at Single-molecule and Single-cell levels

arxiv: v2 [cond-mat.stat-mech] 18 Jun 2009

Non equilibrium thermodynamics: foundations, scope, and extension to the meso scale. Miguel Rubi

Fluctuations in Small Systems the case of single molecule experiments

A Brownian ratchet driven by Coulomb friction

Nonequilibrium Thermodynamics of Small Systems: Classical and Quantum Aspects. Massimiliano Esposito

arxiv:physics/ v2 [physics.class-ph] 18 Dec 2006

arxiv:cond-mat/ v2 [cond-mat.stat-mech] 25 Sep 2000

Large deviation functions and fluctuation theorems in heat transport

Dynamics of two coupled particles: comparison of Lennard-Jones and spring forces

Tightening the uncertainty principle for stochastic currents

Onsager theory: overview

Session 1: Probability and Markov chains

Boundary Dissipation in a Driven Hard Disk System

Smoluchowski Diffusion Equation

arxiv: v3 [cond-mat.stat-mech] 13 Nov 2007

16. Working with the Langevin and Fokker-Planck equations

Anatoly B. Kolomeisky. Department of Chemistry CAN WE UNDERSTAND THE COMPLEX DYNAMICS OF MOTOR PROTEINS USING SIMPLE STOCHASTIC MODELS?

LANGEVIN THEORY OF BROWNIAN MOTION. Contents. 1 Langevin theory. 1 Langevin theory 1. 2 The Ornstein-Uhlenbeck process 8

The Jarzynski Equation and the Fluctuation Theorem

STOCHASTIC PROCESSES IN PHYSICS AND CHEMISTRY

arxiv: v1 [cond-mat.stat-mech] 7 Mar 2019

Power-law behaviors from the two-variable Langevin equation: Ito s and Stratonovich s Fokker-Planck equations. Guo Ran, Du Jiulin *

Anomalous diffusion in biology: fractional Brownian motion, Lévy flights

From time series to superstatistics

Entropy and Free Energy in Biology

The energy speed accuracy trade-off in sensory adaptation

Dissipation and the Relaxation to Equilibrium

Théorie des grandes déviations: Des mathématiques à la physique

Thermodynamic time asymmetry in nonequilibrium fluctuations

Emergent Fluctuation Theorem for Pure Quantum States

Lecture 3: From Random Walks to Continuum Diffusion

4. The Green Kubo Relations

Superstatistics and temperature fluctuations. F. Sattin 1. Padova, Italy

Confined Brownian Motion:

arxiv: v1 [cond-mat.stat-mech] 18 May 2012

Active Matter Lectures for the 2011 ICTP School on Mathematics and Physics of Soft and Biological Matter Lecture 3: Hydrodynamics of SP Hard Rods

Irreversibility and the arrow of time in a quenched quantum system. Eric Lutz Department of Physics University of Erlangen-Nuremberg

For slowly varying probabilities, the continuum form of these equations is. = (r + d)p T (x) (u + l)p D (x) ar x p T(x, t) + a2 r

Numerical study of the steady state fluctuation relations far from equilibrium

PROBABILITY: LIMIT THEOREMS II, SPRING HOMEWORK PROBLEMS

arxiv: v1 [cond-mat.stat-mech] 3 Apr 2007

Introduction to nonequilibrium physics

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 6 Feb 2004

Transcription:

Infimum Law and First-Passage-Time Fluctuation Theorem for Entropy Production arxiv:64.459v [cond-mat.stat-mech] 4 Apr 26 Izaak Neri,, 2 Édgar Roldán, and Frank Jülicher Max Planck Institute for the Physics of Complex Systems, Nöthnitzer Str. 38, 87 Dresden, Germany. 2 Max Planck Institute of Molecular Cell Biology and Genetics, Pfotenhauerstraße 8, 37 Dresden, Germany. We derive an infimum law and a first-passage-time fluctuation theorem for entropy production of stochastic processes at steady state. We show that the ratio between the probability densities of the first-passage time to produce S tot of entropy and of the first-passage time to produce S tot of entropy equals e S tot/k, with k Boltzmann s constant. This first-passage-time fluctuation relation is valid for processes with one or higher order passages. In addition, we derive universal bounds for the infimum statistics of entropy production using the fact that at steady state e S tot/k is a martingale process. We show that the mean value of the entropy-production infimum obeys inf S tot k and derive a bound on the cumulative distribution of entropy-production infima. Our results are derived using a measure-theoretic formalism of stochastic thermodynamics. We illustrate our results with a drift-diffusion process and with numerical simulations of a Smoluchowski Feynman ratchet. PACS numbers: 5.4.-a, 5.7.-a, 2.5.-r, 5.7.Ln I. INTRODUCTION AND STATEMENT OF THE MAIN RESULTS The second law of thermodynamics states that the total entropy in a thermodynamic process increases with time. If such entropy production S tot occurs in a mesoscopic system that is in contact with a large environment, S tot is a stochastic quantity and the second law should be considered as a statement on the average entropy production. This idea was already formulated by J C Maxwell in the 9th century [], but only in the last decades it has been formalized using the language of stochastic processes [2]. For stochastic processes it is possible to define the entropy produced along a single realization of the process consistent with thermodynamics [3 6]. The stochastic definition of entropy production implies a set of relations for the entropy-production fluctuations that are called fluctuation theorems [7]. An example is the detailed fluctuation theorem, which can be written as p S (S tot ; t) p S ( S tot ; t) = estot/k, () where k is Boltzmann s constant. Here p S (S tot ; t) is the probability density for the entropy production S tot at a given time t. The detailed fluctuation theorem Eq. () is universal and holds for a broad class of physical processes, such as, deterministic nonequilibrium processes coupled to thermostats [8 ], stochastic Markovian processes [5, 3] and stochastic non-markovian processes [4, 5]. Moreover, the steady-state fluctuation theorem given by Eq. () has been tested in several experiments, inter alia, a Rayleigh-Bénard convection cell [6], a dragged colloidal particle in an optical trap [7], a fluidized steady-state of a granular medium [8], a photochromic defect center in a diamond [9], biological systems such as the F-ATPase rotary motor [2] and a single-electron box [2]. Until now, most work has focused on entropyproduction fluctuations in a fixed time interval [6, 7, 22, 23]. However, little is known about fluctuations of other statistical properties of entropy production [24 26]. Do first-passage-time fluctuations and extreme-value statistics of entropy production satisfy universal equalities? In this paper we give some novel insights on this question. Here we derive universal fluctuation relations for the first-passage-time distributions of entropy production of classical and continuous stochastic processes in steady state. We call these relations first-passage-time fluctuation theorems. The simplest example is p T (t; S tot ) p T (t; S tot ) = estot/k. (2) Here p T (t; S tot ) is the first-passage-time distribution of entropy production; more precisely p T (t; S tot ) is the probability density for the first-passage time t at which the process has produced for the first time the entropy S tot. The first-passage-time fluctuation theorem is illustrated in Fig. A. Very interesting are also the extreme-value statistics of entropy production. We show that the average of the infimum of the stochastic entropy production in a given time interval is bounded from below by k: inf Stot k. (3) This infimum law is illustrated in Fig. B and bears similarity with the second law of thermodynamics S tot. We show that the infimum law follows from a universal lower bound on the cumulative distribution of the infimum of entropy production: ( Pr inf S ) tot k s e s, (4)

2 for s. Here Pr ( ) denotes the probability of an event. Remarkably, as shown here the infimum law given by Eq. (3) is generally valid for classical and stationary stochastic processes. The paper is structured as follows: In Sec. II, we briefly review the formalism of stochastic thermodynamics based on path-probability densities. In Sec. III, we formulate stochastic thermodynamics using measure theory, define entropy production for a stationary stochastic process and discuss its connection to martingale processes. The first-passage-time fluctuation theorem and the infimum law are derived in Sec. IV and Sec. V, respectively. In Sec.VI, we study the drift-diffusion process for which we present exact expressions of first-passage-time distributions and infimum distributions of entropy production. In Sec. VII we study a Smoluchowski-Feynman ratchet for which we obtain numerical estimates of the first-passage time and infimum statistics of entropy production. For both examples the first-passage-time fluctuation theorems and the infimum law are illustrated. We end with a discussion and an outlook in Sec. VIII. II. STOCHASTIC THERMODYNAMICS AND ENTROPY PRODUCTION We briefly review the basic formalism of stochastic thermodynamics and the definition of entropy production based on path-probability densities [7, 27, 28]. We consider the dynamics of a mesoscopic system in a nonequilibrium steady state described by the coarse-grained state variables ω(t) = (q(t), q(t)) consisting of the mesoscopic degrees of freedom q(t) and q(t) at time t. The set q(t) represents n degrees of freedom that are even under time reversal and the set q(t) contains ñ degrees of freedom that are odd under time reversal [29]. The variables q(t) and q(t) could represent the dynamics of collective modes in a system of interacting particles, for instance, q(t) could be the position and q(t) the effective momentum of a colloidal particle in a fluid. In a given time window [, t] the coordinates ω(t) trace a path in phase space ω t = {q(τ), q(τ)} τ t. We associate with each trajectory ω t a probability density P(ω) t that captures the limited information provided by the coarse-grained variables ω and the fact that the exact microstate is not known. The entropy production (or total entropy change) associated with a path ω t in a stationary process is given by [3, 3] S tot (t) = k ln P (ωt ) P (Θ t ω t ), (5) where Θ t ω t = {q(t τ), q(t τ)} t τ= is the timereversed trajectory. The steady-state average entropy production during a time interval t can be expressed as a path integral [27, 32]: S tot (t) = k Dω t P(ω) t ln P(ωt ) P(Θω t). (6) A. First-passage-time Fluctuation Theorem Entropy production B. The Infimum Law Entropy production Time Time FIG. : Illustration of the two theorems derived in the paper. A. First-passage-time fluctuation theorem for entropy production for a process with two absorbing boundaries. Examples of trajectories of the stochastic entropy production as a function of time that first reach a positive threshold S tot (blue thick curves) and first reach a negative threshold S tot (red thin curves). The probability distributions to first reach the positive p T (t; S tot) and the negative p T (t; S tot) thresholds at time t are related by Eq. (2). B. Infimum law for entropy production. The infima of trajectory samples of total entropy production are shown in dashed lines. The mean of the infimum of the entropy production in a time interval t (teal solid line) is greater than k (yellow thick solid line). Entropy production is therefore the observable that quantifies time irreversibility of mesoscopic trajectories [33]. In fact by measuring entropy production an observer can determine within a minimal time whether a movie of a stochastic process is run forwards or backwards [24]. Microscopic reversibility implies local detailed balance of the conditional probability densities which reads [34, 35]: P (ω t ω()) P (Θ t ω t ω(t)) = esenv(t)/k, (7) where S env (t) is the entropy change in the environment. The total entropy change S tot (t) can be written as S tot (t) = S sys (t) + S env (t), (8)

3 with S sys (t) = k ln P (ω(t)) P (ω()). (9) We call S sys (t) the change in system s entropy [5]. For a system in contact with one or several thermal baths the environment entropy change is related to the heat exchanged between system and environment [36]. An important property of entropy production is that its exponential e Stot(t)/k is a martingale process for stationary processes. A process is martingale when its expected value at any time t equals to its value at a previous time τ, when the expected value is conditioned on observations up to the time τ. In other words, the expectation value of the exponential of minus the entropy change at time t equals its value at the time τ, when the expectation value is conditioned on the trajectory ω τ [37]: e Stot(t)/k ω τ = e Stot(τ)/k. () In addition, as discussed in Appendix B, a martingale process has to be integrable. Note that the martingale property of e Stot(t)/k implies Jarzynski s equality e Stot(t)/k = [38, 39]. The martingale property of e Stot(t)/k is key to derive our two main results in sections IV and V. III. MEASURE-THEORETIC FORMALISM OF STOCHASTIC THERMODYNAMICS In discrete time, the expressions Eqs. (5) and (6) for entropy production are well-defined [4]. In continuous time, these expressions are problematic since the path-probability densities P are not well defined in the continuous-time limit [4]. We therefore use a more general mathematical formalism based on measure theory. This measure-theoretic formalism of stochastic thermodynamics is presented in this section. We first review fundamental concepts of measure theory [42 46]. We then use these concepts to define the entropy production of a continuous stochastic process and discuss its properties. A. Probability space of a stochastic process We introduce the concept of a probability measure P as a generalization of the path probability densities P commonly used in stochastic thermodynamics. The difference between probability densities and probability measures is that a density associates a weight to one trajectory ω while a measure associates a weight to sets of trajectories Φ. The value P (Φ) denotes the probability to observe a trajectory ω in the set Φ, in other words P (Φ) = Pr (ω Φ). In measure theory the set Ω is called the sample space and the set F of all measurable sets Φ is a σ-algebra. The triple (Ω, F, P) is called a probability space. Note that here and below ω = {q(τ), q(τ)} τ (, ) refers to a full trajectory of the state variables. An important example of a measure P on a sample space in R n is given by P (Φ) = dx p(x), () Φ where we have used the Lebesgue integral and p(x) is the probability density of the measure P. It is also possible to define a probability density of a measure P(Φ) = dp with respect to a suitable reference measure Q using the Radon-Nikodým theorem: P (Φ) = dq R(ω), (2) where we have introduced an integral over a probability space as discussed in Appendix A. The function R(ω) is called the Radon-Nikodým derivative, which we usually write as R(ω) = dp dq (ω). In Eq. (2) R(ω) is a generalization of the probability density p(x) to spaces for which the Lebesgue measure does not exist, e.g., the Wiener space of trajectories of a Brownian particle (see Appendix A). A stochastic process X(ω; t) provides the value of an observable X at time t for a given trajectory ω. The average or expectation value of the stochastic variable X(t) is denoted by X(t) P = X(ω; t) dp. We often ω Ω write simply X(t) and X(t) to refer to the stochastic process and its average. In Appendix B we present formal definitions of stochastic processes. In what follows, we focus on stationary probability measures of steady-state processes. A stationary measure is time-translation invariant, i.e. it satisfies P = P T t for all values of t. Time translation T t is a map on trajectory space with the property X(T t (ω); τ) = X(ω; τ + t) for any stochastic process X. B. General definition of stochastic entropy production We now define the stochastic entropy production as a stochastic process on a stationary probability measure [3], which is also discussed in Appendix C. We express the entropy production in a time interval [, t] using the Radon-Nikodým derivative of the measure P F(t) with respect to the time-reversed measure (P Θ) F(t), viz., S tot (ω; t) = k ln dp F(t) d (P Θ) F(t) (ω). (3) In Eq.(3) we have introduced the restricted measure P F(t) over those events in the sub-σ-algebra F(t) F that correspond to finite-time trajectories ω t in the time window [, t] (see Fig. 2A and Appendix B). Here the time-reversed measure P Θ is defined in terms of the

4 A A. One absorbing boundary B FIG. 2: A. Illustration of trajectories of a coarse-grained variable X in a time window [, τ]. Here F(τ) is the sub-σ-algebra generated by trajectories in the time window [, τ] and P F(τ) is the restricted measure over this time window. B. Illustration of the time-reversal map Θ on trajectories ω, ω 2 and ω 3 of the coarse-grained variable X. B. Two absorbing boundaries time-reversal map Θ, which acts on trajectories and maps future events on past events (see Fig. 2B). Time reversal Θ is defined by the property X(Θ(ω); t) = X(ω; t) for all X that are even under time reversal and X(Θ(ω); t) = X(ω; t) for all X that are odd under time reversal. For stochastic processes in continuous time, such as continuous processes or processes with jumps, Eq. (3) provides the entropy production in a given time interval [, t]. In discrete time, Eq. (3) becomes the ratio of the path probability densities corresponding to the forward and backward dynamics and is equal to Eq. (5). Note that for systems that are microreversible Eq. (3) always exists [47]. The martingale property of e Stot(t)/k (as well as e Stot(t)/k ) holds for the general definition of entropy production of stationary process given by Eq. (3). As shown in Appendix C this follows directly from the fact that e Stot(t)/k and e Stot(t)/k are Radon-Nikodým density processes. Using the theory of martingales developed by Doob [42, 43] we derive two new general relations for entropy production. In the next section we derive the firstpassage-time fluctuation theorem and in the subsequent section the infimum law for entropy production. IV. FIRST-PASSAGE-TIME FLUCTUATION THEOREMS In this section we derive the first-passage-time fluctuation theorems. We first introduce stopping times that generalize passage times. We then derive the first- FIG. 3: Illustration of the first-passage times for entropy production. A. First-passage time T () for entropy production with one positive absorbing boundary s tot (top) and firstpassage time T () for entropy production with one negative absorbing boundary s tot (bottom). B. First-passage times T (2) (2) and T for entropy production with two absorbing boundaries at ±s tot. At the time T (2) entropy production passes for the first time the threshold s tot without having (2) reached s tot before. At the time T entropy production passes for the first time the threshold s tot without having reached s tot before. passage theorems and present the main implications on entropy production fluctuations. Formal definitions of stopping times, passage times and the derivation of the first-passage-time fluctuation theorem can be found in Appendix D. A. General form of the first-passage-time fluctuation theorem We now consider processes that start at the reference time t = and that are terminated at a random time t = T, which is called the stopping time. The stopping time can depend on a criterion that the observables of the process have to satisfy in order to terminate the process. We define a s tot -stopping time T based on entropy production such that S tot (T ) = s tot when the process

5 terminates. An example of such an s tot -stopping time is the firstpassage time T (), which determines the time at which the entropy change S tot (t) reaches for the first time the value s tot. Another example is given by the first-passage time T (2), the time at which entropy S tot(t) passes for the first time a threshold value s tot, given that it has not reached s tot before. This implies that if the process first (2) reaches s tot it terminates at a time T. This process is therefore equivalent to a first-passage problem with two absorbing boundaries. Note that there exist many other s tot -stopping times such as times of second passage. The first-passage times T () (2) and T are illustrated in Fig. 3. We have derived the following general fluctuation theorem for s tot -stopping times T in Appendix D: P (Φ T t ) P (Θ T [Φ T t ]) = estot/k, (4) where P (Φ T t ) is the probability to observe a trajectory in the set Φ T t of infinite-time trajectories ω that satisfy the termination criterion at a time T t. Note that these sets contain trajectories that do not actually terminate. The set Θ T [Φ T t ] describes the time-reversed trajectories of Φ T t. It is generated by applying the time-reversal map Θ T to all the elements of the original set. The map Θ T = T T Θ corresponds to a time reversal within a time window [, T ] with the property that X(Θ T (ω); τ) = X(ω; T (ω) τ) for all stochastic processes X that are even under time reversal. The fluctuation theorem for s tot -stopping times Eq. (4) is valid for continuous and stationary stochastic processes. In our derivation we use the martingale property of e Stot(t)/k. The probability density p T of the s tot -stopping time T is given by: p T (t; s tot ) = d dt P (Φ T t). (5) Entropy production is odd under time reversal S tot (Θ T (ω); T (ω)) = S tot (ω; T (ω)) as shown in Appendix C. Therefore, the first-passage times T = T () and T = T (2) have the property T () Θ T (Φ T t ) = Φ T t, (6) T (2) where T = and T = are the corresponding firstpassage times for the threshold s tot illustrated in Fig. 3. From this follows p T (t; s tot ) = d dt P (Θ T [Φ T t ]). (7) For all s tot -stopping times for which T defined in Eq. (6) is the stopping time for the threshold s tot Eq. (7) holds and we also have p T (t; s tot ) p T (t; s tot ) = estot/k. (8) The fluctuation theorem given by Eq. (8) therefore not only holds for distributions of T () (2) and T but also for higher order passage processes that depend on a single threshold s tot. It generalizes the results derived in Refs. [24] and [25] to stationary and continuous stochastic processes. In [24] a first-passage-time fluctuation relation analogous to Eq. (8) has been derived for stochastic processes with translation-invariant transition rates. In Ref. [25] the authors find the fluctuation relation given by Eq. (8) in the asymptotic limit t. Equation (8) implies two interesting results for first-passage times of stochastic processes that are outlined below. B. Fluctuation theorem for passage probabilities The first-passage-time fluctuation theorem implies a fluctuation relation between passage probabilities of entropy production: P + P = e stot/k. (9) Passage probabilities are the probabilities for the process to terminate at a boundary. We distinguish two processes, one which terminates at s tot and one which terminates at s tot, with passage probabilities P + = P = dt p T (t; s tot ), (2) dt p T (t; s tot ), (2) and with P ±. Equation (9) follows directly from integrating Eq. (8) over time. For the case of the first-passage times T (2) with two absorbing boundaries the process terminates in a finite time either in the positive or the negative threshold with probability one and therefore P + + P =. In this case, we find the following exact expressions for the two passage probabilities in terms of the threshold value s tot : estot/k P + = + e stot/k, (22) P = + e stot/k. (23) C. Symmetry of the normalized first-passage-time distributions The first-passage-time fluctuation relation Eq. (8) also implies an equality between the normalized firstpassage-time distributions p T (t s tot ) and p T (t s tot ) which reads p T (t s tot ) = p T (t s tot ). (24)

6 The normalized distributions are defined as: p T (t s tot ) = p T (t; s tot ) dt p T (t; s tot ), (25) p T (t s tot ) = p T (t; s tot ) dt p T (t; s tot ). (26) The symmetric relation Eq. (24) comes from the fact that the ratio of the first-passage-time distributions in Eq. (8) is time independent. The first-passage-time fluctuation theorem thus implies that the mean first-passage time given that the process terminates at the positive boundary, equals to the mean first-passage time given that the process terminates at the negative boundary. This remarkable symmetry extends to all the moments of the first-passagetime distributions. A similar result has been found for waiting-time distributions in chemical kinetics [48 54] for the cycle-time distributions in Markov chains [53, 55] and for decision-time distributions in sequential hypothesis tests [56]. These results could therefore be interpreted as a consequence of the fundamental relation for the firstpassage time fluctuations of entropy production given by Eq. (24). V. THE INFIMUM LAW We now derive the infimum law valid for nonequilibrium steady-state processes. We only need the martingale property of e Stot(t)/k with respect to P. The cumulative distribution of the the supremum of e Stot(t)/k satisfies { Pr (sup τ [,t] e Stot(τ)/k} ) λ e Stot(t)/k, λ (27) with λ [26]. Equation (27) corresponds to Doob s maximal inequality for martingale processes given by Eq. (B5) in Appendix B. Using Jarzynski s equality e Stot(t)/k = [38], Eq. (27) implies a lower bound on the cumulative distribution of the infimum of S tot in a given time interval [, t]: ( ) inf Stot (t) Pr s e s, (28) k with s and inf S tot (t) = inf τ [,t] {S tot (τ)}. The right hand side of Eq. (28) is the cumulative distribution of an exponential random variable S with distribution function p S (s) = e s. From Eq. (28) it thus follows that the random variable inf S tot (t)/k dominates stochastically over S. Stochastic dominance implies an inequality on the mean values of the corresponding random variables as shown in Appendix E. Equation (28) therefore.8.6.4.2 2 - FIG. 4: Sketch of a drift-diffusion process with periodic boundary conditions. In this example, a Brownian particle (gray sphere) with diffusion coefficient D is confined on a ring of radius R = 2 and is drifted with mean velocity v. Here V (φ) is the potential felt by the particle and φ its azimuthal angle. implies the following universal bound for the mean infimum of entropy production: -2-2 inf S tot (t) k. (29) The infimum law given by Eq. (29) holds for stationary stochastic processes in discrete time and for stationary stochastic processes in continuous time for which e Stot(t)/k is càdlàg [57]. Càdlàg processes are stochastic processes that may contain jumps as explained in Appendix B. VI. DRIFT-DIFFUSION PROCESS As a first illustrative example we study a drift-diffusion process of a Brownian particle with diffusion coefficient D, average drift velocity v, and periodic boundary conditions with period l (see Fig. 4). For simplicity we consider here the case of a one-dimensional Brownian motion without inertia. The state of the particle at time t can be described by a variable φ(t) [, l). In the illustration of a ring geometry in Fig. 4, φ is the azimuthal angle and l = 2π. Equivalently, one can consider a stochastic process X(t) given by the net distance traveled by the Brownian particle up to time t: X(t) = φ(t) + ln φ (t), where N φ (t) is the winding number or the net number of clockwise turns (or minus the number of counterclockwise turns) done by the particle up to time t [25]. The time evolution of X(t) is described by the following Langevin equation dx(t) dt - = v + ζ(t). (3) The term ζ(t) in Eq. (3) is a Gaussian white noise with zero-mean ζ(t) = and with autocorrelation 2

7 ζ(t)ζ(t ) = 2Dδ(t t ). The entropy production in a time t in steady state is given by [24] S tot (t) = k v [X(t) X()], (3) D or equivalently S tot (t) = k v D [φ(t) φ()] + kl v D N φ(t). Equation (3) implies that the first-passage and extreme value statistics of entropy production in the driftdiffusion process can be obtained from the statistics of the position X(t) of a drifted Brownian particle in the real line. The drift-diffusion process is an example for which the first-passage-time fluctuation theorem and the infimum law can be verified analytically, as we show below. A. First-passage-time fluctuation theorem The first-passage-time distribution for X to pass at time t for the first time the threshold L > starting from the initial condition X() = is given by Wald s distribution [58, 59] p T (t; L) = Equation (32) implies: L 4πDt 3 e (L vt)2 /4Dt. (32) p T (t; L) p T (t; L) = evl/d. (33) Note that the argument of the exponential equals to the Péclet number, Pe = vl/d. The distribution of the entropy first-passage time T () equals to the first-passage-time distribution given by Eq. (32) for the position of the particle with an absorbing boundary at L = s tot D/vk. Replacing L by s tot D/vk in Eq. (33) one finds the first-passage-time fluctuation theorem for entropy production p T (t; s tot ) p T (t; s tot ) = estot/k. (34) An analogous relation holds for the two-boundary firstpassage times T (2) for entropy production and can be derived using the results in Sec. 2.2.2.2 in [58] (see also [24]). B. Infimum law For the drift-diffusion process in steady state we derive an exact expression for the cumulative distribution of minus the infimum of entropy production (see Appendix F): ( Pr inf S ) tot(t) s (35) k [ ( ) ( )] = s σ(t) erfc 2 2 e s s σ(t) erfc σ(t) 2, σ(t) Cumulative distribution.8.6.4.2 Bound -2 - FIG. 5: Cumulative density distribution of the infimum entropy production in a drift-diffusion process, for different values of the mean entropy change σ(t) = S tot(t) /k. The different curves are calculated using Eq. (35) and the bound given by e s is also shown (yellow thick line). where s >, erfc is the complementary error function and σ(t) = S tot (t) /k = (v 2 /D)t is the average entropy production in steady state at time t. Equation (35) illustrates the universal bound on the infimum cumulative distribution given by Eq. (28). Indeed, in Fig. 5 we show that the infimum cumulative distribution ( given by ) Eq. (35) is bounded from below by Pr inf Stot(t) k s e s, for different values of σ(t). Interestingly, the lower bound is reached when σ(t) becomes very large, which corresponds to the asymptotic limit of long times or the very irreversible limit at finite times t. ( In the reversible) limit of small σ(t), the probability Pr inf Stot(t) k s, which results in a trivial upper bound for the cumulative distribution. We also derive the following analytical expression for the mean infimum of the entropy production in steady state (see Appendix F): inf S tot(t) (36) k [ ] [ ] σ(t) = erf + σ(t) σ(t) σ(t) 2 2 erfc 2 π e σ(t)/4, where erf is the error function. Equation (36) satisfies the infimum law inf S tot (t) k. Interestingly the lower bound for the mean infimum is reached in the limit of large mean entropy production σ(t), as shown in Fig 6. In the equilibrium limit we have inf S tot (t). VII. SMOLUCHOWSKI FEYNMAN RATCHET We now discuss the statistics of first-passage times and infima of entropy production in a paradigmatic ex-

8 -.2 -.4 -.6 -.8 - -2-2 FIG. 6: Mean value of the entropy production infimum given by Eq. (36) as a function of the mean entropy change σ(t) = S tot(t) /k (blue thin curve). The lower bound of the mean infimum set by the infimum law in k is also shown (horizontal yellow thick line). ample of a nonequilibrium process for which system entropy is varies with position in steady state. We study a Smoluchowski-Feynman ratchet for which an overdamped Brownian particle moves under influence of an constant external force in a periodic potential [6, 6] (see Fig. 7 for a graphical illustration). Experimentally this model has been realized using colloidal particles trapped with toroidal optical potentials [62 65]. The steady-state fluctuation theorem given by Eq. () was tested using this experimental technique in Ref. [62]. Analogously to the drift-diffusion process, we describe the dynamics of the Smoluchowski Feynman ratchet in terms of the net distance X(t) traveled by the particle. The dynamics of the stochastic process X(t) is given by an overdamped Langevin equation γ dx(t) dt V (X(t)) = + F + ξ(t), (37) x where γ is a friction coefficient and ξ is a Gaussian white noise with zero mean ξ(t) = and with autocorrelation ξ(t)ξ(t ) = 2kT γ δ(t t ). Here V (x) is a periodic potential of period l, V (x + ml) = V (x) with m Z, and V (X(t)) x = V (x) x. X(t) The system entropy change in steady state over a time t equals [6, 66] S sys (t) = U(t) T + k ln X()+l X() X(t)+l X(t) dy e U(y)/kT dy e U(y)/kT, (38) where U(t) = U(X(t)) U(X()) with U(x) = V (x) F x equal to the effective potential felt by the particle. The change in the environment entropy is given by S env (t) = Q(t)/T, where Q(t) is the heat transferred.8.6.4.2 2 - - -2-2 FIG. 7: Illustration of a Smoluchowski Feynman ratchet. A Brownian particle (gray sphere) immersed in a thermal bath of temperature T moves in a periodic potential V (φ) (black shaded curve) with friction coefficient γ. The coordinate φ is the azimuthal angle of the particle. When applying an external force F in the azimuthal direction, the particle reaches a nonequilibrium steady state. In this example, V (φ) = kt ln[cos(φ) + 2], α = R cos(φ) and β = R sin(φ), with R = 2. from the environment to the particle in a time t. Following Sekimoto [4], Q(t) = t ( γdx(s)/ds + ξ(s)) dx(s) = t U(X(s))/ x dx(s) = U(t), where denotes the Stratonovich product. Therefore, in this example, S env (t) = U(t) T. (39) The change in total entropy over a time t is given by S tot (t) = S sys (t) + S env (t), with: For the choice S tot (t) = k ln X()+l X() X(t)+l X(t) dy e U(y)/kT dy e U(y)/kT. (4) V (x) = kt ln[cos(2πx/l) + 2] (4) the integrals in Eq. (4) can be calculated analytically [66]. The stochastic entropy production in steady state given by Eq. (4) equals S tot (t) k = f[x(t) X()] ln ψ(x(t), f) ψ(x(), f) 2, (42) with f = F l/2πkt, ψ(x, f) = f 2 [cos(x)+2] f sin(x)+2. We perform numerical simulations of Smoluchowski- Feynman ratchet Eq. (37) in steady state with a potential given by Eq. (4). We then obtain numerical estimates

9 of the first-passage times of entropy production and of the infima distributions of entropy production using the expression for entropy production (42). These empirical estimates are used to test our universal results. A. First-passage-time fluctuation theorems First we study the first-passage times T (2) for entropy production with two absorbing boundaries at the threshold values s tot and s tot (with s tot > ). Figure 8 shows the empirical first-passage-time distribution p T (t; s tot ) to first reach the positive threshold (blue squares) together with the first-passage time distribution p T (t; s tot ) (scaled by a factor e stot/k ) to first reach the negative threshold (red circles). Since both distributions coincide we confirm the validity of the first-passagetime fluctuation theorem given by Eq. (2) for the twoboundary first-passage problem in the Smoluchowski Feynman ratchet. Moreover, the functional dependency of the empirical passage probabilities P + and P on the threshold value s tot obeys the expressions given by Eqs. (22) and (23) (see top inset in Fig. 8). As a result, the integral firstpassage-time fluctuation theorem given by Eq. (9) is also fulfilled in this example (see bottom inset in Fig. 8). Note that a method to calculate analytically the passage probabilities P + and P for one-dimensional Langevin processes with two absorbing boundaries was introduced in Refs. [67 69]. As a second case we consider two one-boundary firstpassage problems for entropy production. We study the first-passage time T () for entropy production with one absorbing boundary at the threshold value s tot and the corresponding first-passage time at the negative threshold value s tot. We obtain numerical estimates of the distribution p T (t; s tot ) for the entropy production to first reach at time t a single positive boundary located at s tot, and its conjugate first-passage-time distribution p T (t; s tot ) for entropy production to reach for the first time the single negative threshold s tot. Figure 9 shows empirical estimates of these first-passage-time distributions and confirms the validity of the first-passage-time fluctuation theorem given by Eq. (8). Note that for small values of the thresholds the passage probabilities P + and P both exceed the values given by the twoboundary case Eqs. (22) and (23) (see top inset in Fig. 9). Nevertheless, the integral first-passage-time fluctuation theorem given by Eq. (9) remains valid as shown in the bottom inset of Fig. 9. B. Infimum law We now study the infimum properties of entropy production in the Smoluchowski Feynman ratchet. First we compute numerical estimates of the cumulative distribution of the entropy production infimum measured Two-boundary first-passage probability - -2-3 -4 2 3 4 Time (s) P +.5 P 2 4 stot/k 4 2 2 4 stot/k FIG. 8: Empirical two-boundary first-passage-time distributions of entropy production to first reach a positive threshold p T (t; s tot) (blue squares) and the rescaled distribution for the negative threshold p T (t; s tot)e stot/k (red circles) for the Smoluchowski-Feynman ratchet in steady state. The distributions are obtained from 4 numerical simulations and the threshold values are set to ±s tot = ±2.2k. The simulations are done using Euler numerical scheme with with the following parameters: F = 4 pn, T = 3 K, l = 2π nm, γ = 8.4 pns/nm and simulation time step t =.27 s. Top inset: Empirical passage probabilities of entropy production to the positive (P +, blue squares) and to the negative (P, red circles) thresholds as a function of the threshold values. The analytical expressions for P + given by Eq. (22) (blue solid line) and P given by Eq. (23) (red dashed line) are also shown. Bottom inset: Logarithm of the ratio between the empirical passage probabilities P + and P as a function of the threshold value (magenta diamonds). The solid line is a line of slope. over a fixed time interval for different values of the external force F. Figure shows that the cumulative distribution of minus the entropy production infimum is bounded from below by e s which confirms the universality of Eq. (28). Interestingly, the bound is tighter for larger values of the average entropy production σ(t) = S tot (t) /k. Strikingly, as shown in Fig. the cumulative distribution for the infimum of entropy production of the Smoluchowski Feynman ratchet is nearly identical to the corresponding cumulative distribution of the drift-diffusion process given by Eq. (35). This equivalence between the infimum cumulative distributions holds even for small values of σ(t) where the shape of the potential V (x) affects the entropy-production fluctuations. In Fig. we furthermore show that the infimum law inf S tot (t) k holds for the Smoluchowski Feynman ratchet using numerical simulations. Moreover, we also find that the mean infimum of entropy production follows the functional dependency on σ(t) given by Eq. (36) obtained for the drift-diffusion process. These results point towards a universal behaviour of the entropy production infimum as a function of the average entropy pro-

One-boundary first-passage probability - -2-3 -4 2 4 stot/k 2 3 4 Time (s).5 4 2 P + P 2 4 stot/k FIG. 9: Empirical one-boundary first-passage-time distributions of entropy production to first reach a positive threshold p T (t; s tot) (blue squares) and the rescaled distribution for the negative threshold p T (t; s tot)e stot/k (red circles) for a Smoluchowski-Feynman ratchet in steady state. The estimate of p T (t; s tot) (p T (t; s tot)) is obtained measuring the time elapsed by the entropy production to first reach a single absorbing boundary in s tot = 2.2k ( s tot = 2.2k) in 4 simulations. The simulations are done with the same parameters as in Fig. 8, and the empirical probabilties are calculated over a total simulation time of τ max = 2 s. Top inset: Empirical passage probabilities of entropy production in the positive-threshold simulations (P +, blue squares) and in the negative-threshold simulations (P, red circles) as a function of the value of the thresholds. The expressions P + given by Eq. (22) (blue solid line) and P given by Eq. (23) (red dashed line) are also shown. Bottom inset: Logarithm of the ratio between P + and P as a function of the threshold value (magenta diamonds). The solid line is a line of slope. duced. Such a universal property has also been found for the first-passage-time distributions of entropy production [25]. Cumulative distribution.8.6.4.2 Bound -2 - FIG. : Cumulative distribution of the infimum of entropy production obtained from numerical simulations of a Smoluchowski Feynman ratchet in steady state for different values of the mean entropy change σ(t) = S tot(t) /k compared with the bound given by e s (solid yellow line). The corresponding dashed curves are the values of the cumulative distribution of the infimum in the drift-diffusion process given by Eq. (35). Simulation parameters: 4 simulations with T = 3 K, l = 2π nm, γ = 8.4 pns/nm, simulation time step t =.27 s and total simulation time t =.4 s. -.2 -.4 -.6 -.8 - Drift-diffusion -2-2 VIII. DISCUSSION AND OUTLOOK In this paper we have derived universal relations for first-passage times and extreme-value statistics for entropy production of stationary stochastic processes. Our results do not follow from the standard fluctuation theorem but from other generic properties of stochastic entropy production in particular its martingale property. We have derived a fluctuation theorem given by Eq. (8) for first-passage times of entropy production valid for continuous stochastic processes in steady state. This theorem states that at any time instance it is exponentially more likely to have at this time first produced a positive amount of entropy as compared to the possibility to have at this time first reduced entropy by the same amount. Secondly, we have discussed properties of the infimum of entropy production. Note that it can only be smaller FIG. : Mean value of the entropy production infimum as a function of the mean entropy change σ(t) = S tot(t) /k obtained from numerical simulations of a Smoluchowski Feynman ratchet in steady state with different simulation time steps t. The yellow thick bar is the bound of infimum law given by k. Simulation parameters: 4 simulations with T = 3 K, l = 2π nm, γ = 8.4 pns/nm, and total simulation time t =.4 s. or equal than zero, since at the initial time entropy is zero.s We have shown that the mean infimum of entropy production in steady state is lower bounded by minus the Boltzmann constant. This bound, which we call the infimum law, holds for stochastic processes in discrete time and stochastic processes in continuous time that may have discontinuous jumps.

The fundamental results derived here imply thermodynamic constraints on mesoscopic nonequilibrium processes that result from microscopic reversibility. For instance, in simple models of enzymatic reactions it has been shown that the waiting-time distributions for the forward and backward reactions are exactly the same [48 52, 54]. Single-molecule data on kinesin motor stepping provides evidence that the waiting-time distributions of forward and backward steps are the same [7]. Interestingly, these relations are very similar to the first-passagetime fluctuation theorem on the normalized distributions given by Eq. (24). This raises the question whether reactions times are first-passage times or entropic stopping times and whether these relations found in chemical kinetics can be understood in terms of entropy-production fluctuations. Analogously, we expect that the infimum law can provide novel bounds on extreme-value statistics in enzymatic reactions and other biomolecular processes. Acknowledgments We thank Prof. em. Dr. sc. techn. Heinrich Meyr, Dr. - Ing. Meik Dörpinghaus and Mostafa Khalili-Marandi for enlightening discussions on decision theory and the Gichman-Skorochod method [67, 68]. We also thank Dr. Andre Cardoso Barato for valuable discussions on stochastic thermodynamics, and Dr. Matteo Polettini for making us aware about the cycle fluctuation theorems in Refs. [53] and [55]. E.R. acknowledges funding from Grupo Interdisciplinar de Sistemas Complejos (GISC, Madrid, Spain) and from Spanish Government, Grant TerMic (FIS24-52486-R). References [] J. C. Maxwell, Taits thermodynamics, Nature, vol. 7, no. 257, pp. 66 67, 878. [2] K. Sekimoto, Stochastic energetics, vol. 799. Springer, Berlin, 2. [3] K. Sekimoto, Kinetic characterization of heat bath and the energetics of thermal ratchet models, Journal of the physical society of Japan, vol. 66, no. 5, pp. 234 237, 997. [4] K. Sekimoto, Langevin equation and thermodynamics, Progress of Theoretical Physics Supplement, vol. 3, pp. 7 27, 998. [5] U. Seifert, Entropy production along a stochastic trajectory and an integral fluctuation theorem, Physical review letters, vol. 95, no. 4, p. 462, 25. [6] C. Jarzynski, Equalities and inequalities: irreversibility and the second law of thermodynamics at the nanoscale, Annu. Rev. Condens. Matter Phys., vol. 2, no., pp. 329 35, 2. [7] U. Seifert, Stochastic thermodynamics, fluctuation theorems and molecular machines, Reports on Progress in Physics, vol. 75, no. 2, p. 26, 22. [8] D. J. Evans, E. Cohen, and G. Morriss, Probability of second law violations in shearing steady states, Physical Review Letters, vol. 7, no. 5, p. 24, 993. [9] D. J. Evans and D. J. Searles, Equilibrium microstates which generate second law violating steady states, Physical Review E, vol. 5, no. 2, p. 645, 994. [] G. Gallavotti and E. Cohen, Dynamical ensembles in nonequilibrium statistical mechanics, Physical Review Letters, vol. 74, no. 4, p. 2694, 995. [] J. Kurchan, Fluctuation theorem for stochastic dynamics, Journal of Physics A: Mathematical and General, vol. 3, no. 6, p. 379, 998. [2] J. L. Lebowitz and H. Spohn, A gallavotti cohen-type symmetry in the large deviation functional for stochastic dynamics, Journal of Statistical Physics, vol. 95, no. -2, pp. 333 365, 999. [3] C. Maes, The fluctuation theorem as a gibbs property, Journal of statistical physics, vol. 95, no. -2, pp. 367 392, 999. [4] T. Speck and U. Seifert, The jarzynski relation, fluctuation theorems, and stochastic thermodynamics for nonmarkovian processes, Journal of Statistical Mechanics: Theory and Experiment, vol. 27, no. 9, p. L92, 27. [5] C. Aron, G. Biroli, and L. F. Cugliandolo, Symmetries of generating functionals of langevin processes with colored multiplicative noise, Journal of Statistical Mechanics: Theory and Experiment, vol. 2, no., p. P8, 2. [6] S. Ciliberto and C. Laroche, An experimental test of the gallavotti-cohen fluctuation theorem, Le Journal de Physique IV, vol. 8, no. PR6, pp. Pr6 25, 998. [7] G. Wang, E. M. Sevick, E. Mittag, D. J. Searles, and D. J. Evans, Experimental demonstration of violations of the second law of thermodynamics for small systems and short time scales, Physical Review Letters, vol. 89, no. 5, p. 56, 22. [8] K. Feitosa and N. Menon, Fluidized granular medium as an instance of the fluctuation theorem, Physical review letters, vol. 92, no. 6, p. 643, 24. [9] C. Tietz, S. Schuler, T. Speck, U. Seifert, and J. Wrachtrup, Measurement of stochastic entropy production, Physical review letters, vol. 97, no. 5, p. 562, 26. [2] K. Hayashi, H. Ueno, R. Iino, and H. Noji, Fluctuation theorem applied to f -atpase, Physical review letters, vol. 4, no. 2, p. 283, 2. [2] O.-P. Saira, Y. Yoon, T. Tanttu, M. Möttönen, D. Averin, and J. P. Pekola, Test of the jarzynski and crooks fluctuation relations in an electronic system, Physical review letters, vol. 9, no. 8, p. 86, 22. [22] P. Pietzonka, A. C. Barato, and U. Seifert, Universal bounds on current fluctuations, arxiv preprint arxiv:52.22, 25. [23] T. R. Gingrich, J. M. Horowitz, N. Perunov, and J. England, Dissipation bounds all steady-state current fluctuations, arxiv preprint arxiv:52.222, 25. [24] É. Roldán, I. Neri, M. Dörpinghaus, H. Meyr, and F. Jülicher, Decision making in the arrow of time,

2 Physical review letters, vol. 5, no. 25, p. 2562, 25. [25] K. Saito and A. Dhar, Waiting for rare entropic fluctuations, arxiv preprint arxiv:54.287, 25. [26] R. Chetrite and S. Gupta, Two refreshing views of fluctuation theorems through kinematics elements and exponential martingale, Journal of Statistical Physics, vol. 43, no. 3, pp. 543 584, 2. [27] C. Van den Broeck and M. Esposito, Ensemble and trajectory thermodynamics: A brief introduction, Physica A: Statistical Mechanics and its Applications, vol. 48, pp. 6 6, 25. [28] E. Roldán and J. M. R. Parrondo, Entropy production and kullback-leibler divergence between stationary trajectories of discrete systems, Phys. Rev. E, vol. 85, p. 329, Mar 22. [29] Variables that are even under time reversal do not change sign under time reversal (e.g. position variables) whereas variables that are odd under time reversal change sign under time reversal (e.g. momenta or current). [3] G. E. Crooks, Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences, Physical Review E, vol. 6, no. 3, p. 272, 999. [3] C. Maes and K. Netočnỳ, Time-reversal and entropy, Journal of statistical physics, vol., no. -2, pp. 269 3, 23. [32] É. Roldán, Irreversibility and dissipation in microscopic systems. Springer, Berlin, 24. [33] C. Maes, On the origin and the use of fluctuation relations for the entropy, Séminaire Poincaré, vol. 2, pp. 29 62, 23. [34] P. G. Bergmann and J. L. Lebowitz, New approach to nonequilibrium processes, Physical Review, vol. 99, no. 2, p. 578, 955. [35] G. E. Crooks, Path-ensemble averages in systems driven far from equilibrium, Physical review E, vol. 6, no. 3, p. 236, 2. [36] For a system in contact with l thermal reservoirs at temperatures T i, i =,..., l, the change of environment entropy S env (t) equals to S env (t) = l i= Qi (t) /Ti, with Q i(t) the heat exchanged between the i-th reservoir and the system. [37] The martingale property follows from Dω t τ P(ωτ t ω τ )e Stot(t)/k = Dωτ t P(ωτ t ω τ ) P(Θ tω t ) = P(ω t ) Dω t τ P(ωτ t ω τ P(Θ ) t ω t ) = P(ωτ t ωτ )P(ωτ ) e S tot(τ)/k. [38] C. Jarzynski, Nonequilibrium equality for free energy differences, Physical Review Letters, vol. 78, no. 4, p. 269, 997. [39] We have e Stot(t)/k = dω()p(ω()) Dω t P(ω ω())e t S tot(t)/k = e Stot()/k =, using Eq. () for τ =. [4] H. Risken, The Fokker-planck equation. Springer, Berlin, 984. [4] The normalization constant of the path probability density diverges in the continuum limit. [42] J. L. Doob, Stochastic processes. John Wiley & Sons, Inc., Chapman & Hall, New York, 953. [43] J. L. Doob, Measure theory, vol. 43. Springer Science & Business Media, Berlin, 994. [44] H. L. Royden and P. Fitzpatrick, Real analysis, vol. 98. Macmillan, New York, 988. [45] Z. Schuss, Theory and applications of stochastic processes: an analytical approach, vol. 7. Springer Science & Business Media, Berlin, 29. [46] T. Tao, An introduction to measure theory, vol. 26. American Mathematical Soc., Providence, Rhode Island, 2. [47] The Radon-Nikodým derivative in Eq. (3) is well defined as long as the measure P F(t) is absolutely continuous with respect to the time-reversed measure P Θ F(t). Absolute continuity means in the present context that irreversible transitions form a set of zero measure. [48] H. Qian and X. S. Xie, Generalized Haldane equation and fluctuation theorem in the steady-state cycle kinetics of single enzymes, Physical Review E, vol. 74, no., p. 92, 26. [49] H. Wang and H. Qian, On detailed balance and reversibility of semi-markov processes and single-molecule enzyme kinetics, Journal of mathematical physics, vol. 48, no., p. 333, 27. [5] A. B. Kolomeisky, E. B. Stukalin, and A. A. Popov, Understanding mechanochemical coupling in kinesins using first-passage-time processes, Phys. Rev. E, vol. 7, p. 392, Mar 25. [5] H. Ge, Waiting cycle times and generalized Haldane equality in the steady-state cycle kinetics of single enzymes, The Journal of Physical Chemistry B, vol. 2, no., pp. 6 7, 28. [52] M. Lindén, Decay times in turnover statistics of single enzymes, Physical review E, vol. 78, no., p. 9, 28. [53] C. Jia, D. Jiang, and M. Qian, Cycle symmetries and circulation fluctuations for discrete-time and continuoustime markov chains, arxiv preprint arxiv:47.263, 24. [54] H. Qian, S. Kjelstrup, A. B. Kolomeisky, and D. Bedeaux, Entropy production in mesoscopic stochastic thermodynamics: Nonequilibrium kinetic cycles driven by chemical potentials, temperatures, and mechanical forces, arxiv preprint arxiv:6.48, 26. [55] M. Bauer and F. Cornu, Affinity and fluctuations in a mesoscopic noria, Journal of Statistical Physics, vol. 55, no. 4, pp. 73 736, 24. [56] M. Dörpinghaus, É. Roldán, I. Neri, H. Meyr, and F. Jülicher, An information theoretic analysis of sequential decision-making, arxiv preprint arxiv:5.8952, 25. [57] Càdlàg is an acronym for continue à droite, limite à gauche, which means everywhere right continuous with left limits. [58] S. Redner, A guide to first-passage processes. Cambridge University Press, Cambridge, England, 2. [59] S. Sato and J. Inoue, Inverse gaussian distribution and its application, Electronics and Communications in Japan (Part III: Fundamental Electronic Science), vol. 77, no., pp. 32 42, 994. [6] P. Reimann, Brownian motors: noisy transport far from equilibrium, Physics reports, vol. 36, no. 2, pp. 57 265, 22. [6] P. Hänggi and F. Marchesoni, Artificial brownian motors: Controlling transport on the nanoscale, Reviews of Modern Physics, vol. 8, no., p. 387, 29. [62] T. Speck, V. Blickle, C. Bechinger, and U. Seifert, Distribution of entropy production for a colloidal particle in a nonequilibrium steady state, EPL (Europhysics Letters), vol. 79, no. 3, p. 32, 27.